删除或更新信息,请邮件至freekaoyan#163.com(#换成@)

Solving second-order nonlinear evolution partial differential equations using deep learning

本站小编 Free考研考试/2022-01-02

<script type="text/javascript" src="https://cdn.bootcss.com/mathjax/2.7.2-beta.0/MathJax.js?config=TeX-AMS-MML_HTMLorMML"></script> <script type='text/x-mathjax-config'> MathJax.Hub.Config({ extensions: ["tex2jax.js"], jax: ["input/TeX", "output/HTML-CSS"], tex2jax: {inlineMath: [ ['$','$'], ["\\(","\\)"] ],displayMath: [ ['$$','$$'], ["\\[","\\]"] ],processEscapes: true}, "HTML-CSS": { availableFonts: ["TeX"] }, "HTML-CSS": {linebreaks: {automatic: true}}, SVG: {linebreaks: {automatic: true}} }); </script> Jun Li(李军)1, Yong Chen(陈勇),2,3,41Shanghai Key Laboratory of Trustworthy Computing, East China Normal University, Shanghai, 200062, China
2School of Mathematical Sciences, Shanghai Key Laboratory of PMMP, Shanghai Key Laboratory of Trustworthy Computing, East China Normal University, Shanghai, 200062, China
3College of Mathematics and Systems Science, Shandong University of Science and Technology, Qingdao, 266590, China
4Department of Physics, Zhejiang Normal University, Jinhua, 321004, China

Received:2020-03-17Revised:2020-05-11Accepted:2020-05-27Online:2020-10-02
Fund supported:National Natural Science Foundation of China.11675054
Shanghai Collaborative Innovation Center of Trustworthy Software for Internet of Things.ZF1213
Science and Technology Commission of Shanghai Municipality.18dz2271000


Abstract
Solving nonlinear evolution partial differential equations has been a longstanding computational challenge. In this paper, we present a universal paradigm of learning the system and extracting patterns from data generated from experiments. Specifically, this framework approximates the latent solution with a deep neural network, which is trained with the constraint of underlying physical laws usually expressed by some equations. In particular, we test the effectiveness of the approach for the Burgers' equation used as an example of second-order nonlinear evolution equations under different initial and boundary conditions. The results also indicate that for soliton solutions, the model training costs significantly less time than other initial conditions.
Keywords: deep learning;nonlinear evolution equations;data-driven solutions;solitons;nonlinear dynamics


PDF (2003KB)MetadataMetricsRelated articlesExportEndNote|Ris|BibtexFavorite
Cite this article
Jun Li(李军), Yong Chen(陈勇). Solving second-order nonlinear evolution partial differential equations using deep learning*. Communications in Theoretical Physics, 2020, 72(10): 105005- doi:10.1088/1572-9494/aba243

1. Introduction

Deep learning has garnered remarkable advances across diverse areas, including computer vision, speech recognition, language translation, natural language understanding, and other tasks [1], with the recent growth of big data and computing resources. It represents the data using multi-layer neural networks, i.e. deep neural networks. Despite the remarkable success in these and related areas, deep learning has not yet been widely used in the field of scientific computing. Moreover, its use in solving nonlinear evolution partial differential equations (PDEs) has emerged more recently [2-5].

PDEs have important applications in physics, engineering, biology, finance and other areas. Solving these equations has still been a computational challenge because many of these equations are difficult to be solved analytically or their analytical solution does not exist. Accordingly, solving them with neural networks as nonlinear approximators [6-10] is a very natural idea, and has been considered in several different forms previously [14-18]. The solution via deep learning method is closed analytic, differentiable and easy to be used in subsequent calculations compared with some traditional numerical approaches. In addition, the deep learning method does not require the discretization of spatial and temporal domains compared with other traditional numerical approaches.

Specifically, in this work, we solve nonlinear evolution equations by approximating the unknown solution with a deep neural network [26-28]. The network is trained to satisfy the equation and corresponding initial-boundary conditions. That is to say, with the help of the automatic differentiation techniques [11], the equation is embedded into the loss function in order to utilize the underlying laws of physics to discover patterns from experimental data. Moreover, this method requires much less data than ones proposed in other previous works [12].

Known that a specific setting that yields impressive results for certain equations could fail for some others, we just study second-order nonlinear evolution PDEs in one time and one space variable; that is, we consider the equations where each contains the dissipative term in addition to other partial derivatives. This family of equations has gained its importance because of applications across a range of scientific areas. Specifically, the equations that will be discussed in this paper, is of the form$\begin{eqnarray}{u}_{t}-{u}_{{xx}}=P(u),\end{eqnarray}$where the solution u is a function of the space variable x and the time variable t, and the term P(u) is a function of u and its derivatives. When $P(u)=-2{{uu}}_{x}$, we obtain the well-known Burgers' equation as the example in this paper. In addition, we will get the famous Fisher equation [13] when $P(u)\,=u(1-u)$.

The paper is organized as follows. In section 2, we introduce the neural network architecture and briefly present some problem setups. In section 3, we consider a simple initial condition of the Burgers' equation to illustrate the algorithm. Then we give our main focus on the soliton phenomena of the Burgers' equation in section 4. In section 5, we demonstrate the capabilities of the algorithm for more general initial conditions of the Burgers' equation. Finally, we conclude the paper in section 6.

2. Method

Although the existence of similar ideas for constraining neural network frameworks using some underlying physical laws (see, e.g. [14]), we revisit them using some more advanced tools, and then apply them to more challenging problems described by nonlinear evolution equations. Specifically, in this method, we approximate the solution u with a deep neural network and accordingly define a residual network to be given by$\begin{eqnarray}f:= {u}_{t}-{ \mathcal N }(u,{u}_{x},{u}_{{xx}}),\end{eqnarray}$where ${ \mathcal N }$ is a nonlinear function. The Burgers' equation corresponds to the case where ${ \mathcal N }={u}_{{xx}}-2{{uu}}_{x}$. Reference [26] also made more detailed analysis when including time t or space x, or both in ${ \mathcal N }$. By the way, ${ \mathcal N }$ can also be regarded as a nonlinear differential operator.

We obtain the derivatives of the network u with respect to time t and space x using automatic differentiation with the aid of Tensorflow [19], which is a very popular and open-source deep learning software library.

Our main goal is to minimize the loss function$\begin{eqnarray}L=\displaystyle \frac{1}{{N}_{u}}\displaystyle \sum _{i=1}^{{N}_{u}}| u({t}_{u}^{i},{x}_{u}^{i})-{u}^{i}{| }^{2}+\displaystyle \frac{1}{{N}_{f}}\displaystyle \sum _{j=1}^{{N}_{f}}| f({t}_{f}^{j},{x}_{f}^{j}){| }^{2},\end{eqnarray}$where $\{{t}_{u}^{i},{x}_{u}^{i},{u}^{i}\}{}_{i=1}^{{N}_{u}}$ denote the initial-boundary training data on the solution network u and $\{{t}_{f}^{j},{x}_{f}^{j}\}{}_{j=1}^{{N}_{f}}$ specify the collocation points for the network f. The first term of the right hand side of equation (3) attempts to fit the solution data and the second term learns to satisfy the residual f. Much research has been done on the convergence of the loss function (see, for example, [21]). In addition, whether there are other candidate objective functions or not will be the future research. In this paper, we just choose to optimize the loss function(3) using the L-BFGS algorithm. In addition, for larger datasets, more efficient algorithms can be readily employed, such as the stochastic gradient descent and Adam [20].

Throughout this work, we use relatively simple multi-layer perceptrons with the Xavier initialization and the hyperbolic tangent ($\tanh $) activation function. Commonly used initializations also include Gaussian initialization and He initialization; commonly used activations also include the sigmoid (σ) function and the rectified linear unit. In addition, we have tried some regularizations, such as batch normalization and dropout, it could not improve the expressiveness of the model and the performance sometimes becomes even worse. Therefore, we do not use additional regularizations except for the underlying laws expressed by the given equation in this paper. More detailed analysis will be conducted in future work.

In the following of this paper, we consider the (1+1)-dimensional Burgers' equation along with Dirichlet boundary conditions (other common boundary conditions also including Neumann boundary conditions and mixed boundary conditions):$\begin{eqnarray}\left\{\begin{array}{l}{u}_{t}+2{{uu}}_{x}-{u}_{{xx}}=0,x\in [-20,20],t\in [-10,10],\\ u({t}_{0},x)={u}_{0}(x),\\ u(t,-20)=a,u(t,20)=b,\end{array}\right.\end{eqnarray}$where ${u}_{0}(x)$ is certain initial function, and a, b are constants.

Now, define $f(t,x)$ to be given by$\begin{eqnarray}f:= {u}_{t}+2{{uu}}_{x}-{u}_{{xx}}.\end{eqnarray}$

To obtain the training and testing datasets, we simulate equation (4) using conventional spectral methods (for some kink-type soliton solutions, we just use certain PDE solvers in Matlab). Specifically, starting from an initial condition ${u}_{0}(x),x\in [-20,20]$ and assuming Dirichlet boundary conditions, we use the Chebfun package [22] with a Fourier discretization with 512 modes and a 4th-order explicit Runge-Kutta integrator with time-step size 1 × 10−04, and then integrate the equation up to the final time t=10. The solution is saved every Δt=0.1 to give us a total of 201 snapshots. Out of this dataset, we generate a smaller training subset by randomly subsampling Nu=100 (usually relatively small) initial-boundary data and Nf=10 000 collocation data points generated usually using the Latin hypercube sampling strategy [23].

In this work, we choose the neural network's architectures in a consistent fashion throughout the paper. Specifically, we represent the latent solution by a 13-layer deep neural network with 40 neurons per hidden layer. Furthermore, let ${ \mathcal N }$ be a specific physical prior, which is expressed by an equation rather than a neural network like with that of u [26]. We use $\tanh (x)$ as the activation function only in the hidden layers rather than the output layer. In addition, all experiments are conducted on a MacBook Pro computer with 2.4 GHz Dual-Core Intel Core i5 processor.

3. A simple example: trigonometric function

Let us start with a simple initial and boundary condition in order to highlight the ability of the method. Here, we consider a cosine initial condition and periodic boundary condition of equation (4) which are given by$\begin{eqnarray}\left\{\begin{array}{l}{u}_{0}(x)=\cos \left(\tfrac{\pi (x+10)}{20}\right),\\ u(t,-20)=u(t,20)=0.\end{array}\right.\end{eqnarray}$

We know that the exact solution in this case is analytically available and relatively easy to solve. Figure 1 illustrates our result for the data-driven solution to the Burgers' equation. Specifically, given a set of initial and boundary data, we attempt to learn the latent solution u by training all learnable parameters of the network using the loss function(3). The top panel of figure 1 compares between the exact solution and the predicted spatiotemporal behavior. The resulting prediction error is measured at 9.21 × 10−03 in the relative ${{\mathbb{L}}}_{2}$-norm. Note that this prediction error is much lower than one reported in some previous works where they used other methods such as Gaussian processes [4]. A more detailed assessment of the predicted solution is presented in the bottom panel of figure 1. In particular, we present a comparison between the exact solution and predicted solution at different time points t=−7.5, −2.5, 7.5. Moreover, the algorithm can accurately capture the intrinsic and intricate nonlinear behavior of the Burgers' equation that leads to a shock around t=−2.0 using only a handful of data, while it is notoriously hard to accurately resolve with classical numerical methods and requires a laborious discretization of the equation [32].

Figure 1.

New window|Download| PPT slide
Figure 1.Cosine initial condition. Top: an exact solution to the Burgers' equation is compared to the corresponding solution of the learned PDE. The model correctly captures the dynamics behavior and accurately reproduces the solution with a relative ${{\mathbb{L}}}_{2}$ error of 9.21 × 10−03. Bottom: comparison between the predicted solutions and exact solutions at the three snapshots (corresponding to the vertical lines in the top panel). The model training took about 19 min.


4. Soliton solutions

The soliton phenomena exist widespread in physics, biology, communications and other scientific disciplines. Although some exact and explicit soliton solutions of evolution equations can be obtained by the Darboux transformation, Hirota's direct method, and inverse scattering transform (see, e.g. [25]), these methods are relatively difficult to extend directly to other equations and there exist more unknown types of solitons. Specifically, we study the soliton behaviors of the Burgers' equation here.

4.1. One-soliton solution

First, we consider the one-soliton initial condition:$\begin{eqnarray}{u}_{0}(x)=-\displaystyle \frac{k\exp \left(k(x-10k)\right)}{1+\exp \left(k(x-10k)\right)}.\end{eqnarray}$

Let k=1, then we know that a=0 and $b=-1$.

Figure 2 summarizes our results for one-soliton solution (the soliton is an anti-kink) to the Burgers' equation. The top panel of figure 2 compares between the exact solution and the predicted spatiotemporal behavior. The resulting prediction error is measured at 2.45 × 10−03 in the relative ${{\mathbb{L}}}_{2}$-norm. More detailed assessments of the predicted dynamics are presented in the middle and bottom panels, where the potential v=−ux is observed, of figure 2. In particular, we present a comparison between the exact solutions and predicted solutions at the three different time instants t=−7.5, −2.5, 7.5. The result indicates that the model can accurately capture the single soliton behavior of the Burgers' equation.

Figure 2.

New window|Download| PPT slide
Figure 2.One-soliton solution. Top: an exact one-soliton solution to the Burgers' equation is compared to the solution of the learned PDE (right panel). The system correctly captures the dynamics and accurately reproduces the solution with a relative ${{\mathbb{L}}}_{2}$ error of 2.45 × 10−03. Middle: comparison of the predicted solutions and exact solutions at the three temporal snapshots. Bottom: comparison between the corresponding predicted and exact solutions of the potential. The training process took approximately half a minute.


From figure 3, we can observe the single solitary wave motion better.

Figure 3.

New window|Download| PPT slide
Figure 3.(a) The spatiotemporal behavior of the reconstructed single soliton. (b) The spatiotemporal dynamics of the corresponding potential.


4.2. Two-soliton solutions

Then, we also consider the two-soliton initial condition:$\begin{eqnarray}{u}_{0}(x)=-\displaystyle \frac{{k}_{1}\exp \left({k}_{1}(x-10{k}_{1})\right)+{k}_{2}\exp \left({k}_{2}(x-10{k}_{2})\right)}{1+\exp \left({k}_{1}(x-10{k}_{1})\right)+\exp \left({k}_{2}(x-10{k}_{2})\right)}.\end{eqnarray}$

When k1=1 and k2=−1, we obtain a=1, b=−1.

Figure 4 summarizes our results for the two-soliton solution to the Burgers' equation. The top panel of figure 4 compares between the exact dynamics and the predicted spatiotemporal behavior and the resulting prediction error is measured at 4.23 × 10−03 in the relative ${{\mathbb{L}}}_{2}$-norm. More detailed assessments of the predicted behavior are presented in the middle and bottom panels of figure 4. Moreover, we present a comparison between the exact solution and predicted solution at different time instants t=−7.5, −2.5, 7.5. The algorithm accurately recovers the two-soliton behavior of the Burgers' equation.

Figure 4.

New window|Download| PPT slide
Figure 4.Two-soliton solution. Top: an exact two-soliton solution to the Burgers' equation (left panel) is compared to the corresponding reconstructed solution of the learned PDE. The system correctly captures the nonlinear dynamics and reproduces the solution with a relative ${{\mathbb{L}}}_{2}$ error of 4.23 × 10−03. Middle: comparison between the predicted and exact dynamics at the temporal snapshots. Bottom: comparison between the potential behaviors. The training took approximately 6.5 min.


From figure 5, we can observe the soliton interaction better. In particular, we clearly see that the two solitons with same amplitudes fuse to one single soliton with a different amplitude at certain specific instant.

Figure 5.

New window|Download| PPT slide
Figure 5.(a) The spatiotemporal behavior of the reconstructed two-soliton solution. (b) The nonlinear interaction of the corresponding potential.


We consider another two-soliton solution when k1=1.5 and k2=−1, which results in a=1, b=−1.5.

Figure 6 summarizes our results for another two-soliton solution to the Burgers' equation. The top panel of figure 6 compares between the exact dynamics and the predicted spatiotemporal behavior. The prediction error is measured at 3.50 × 10−03 in the relative ${{\mathbb{L}}}_{2}$-norm. More detailed assessments of the predicted solution are presented in the middle and bottom panels of figure 6. In particular, we present a comparison between the exact solution and predicted solution at different instants t=−7.5, −2.5, 7.5. We find that the algorithm can also accurately capture the two-soliton behavior. From these experiments above, we just observe the soliton fusion behaviors of the Burgers' equation. More detailed analysis have been made, for example in [25]. In addition, the solution degenerates into a simple one-soliton solution whose wave dynamics is like that of section 4.1 when k1=1.5 and k2=1.0.

Figure 6.

New window|Download| PPT slide
Figure 6.Another two-soliton solution. Top: an exact two-soliton solution to the Burgers' equation (left panel) is compared to the solution of the learned PDE. The system correctly captures the nonlinear dynamics and accurately reproduces the solution with a relative ${{\mathbb{L}}}_{2}$ error of 3.50 × 10−03. Middle: comparison between the predicted and exact solutions at the three temporal snapshots. Bottom: comparison between the corresponding predicted and exact solutions of the potential. The model training took about 3.5 min.


From figure 7, we observe that the two single solitons with different amplitudes fuse into one soliton with an amplitude different from the first two again.

Figure 7.

New window|Download| PPT slide
Figure 7.(a) The spatiotemporal behavior of another reconstructed two-soliton solution. (b) The nonlinear interaction of the corresponding potential.


5. More complicated cases

More complicated and even unstable solutions rather than simple and stable solutions (e.g. solitons) often occur in the real world.

5.1. Exponential functions

First, we consider an exponential initial condition and periodic boundary condition of equation (4) given by$\begin{eqnarray}\left\{\begin{array}{l}{u}_{0}(x)=-2\exp \left(-{\left(x+5\right)}^{2}\right),\\ u(t,-20)=u(t,20).\end{array}\right.\end{eqnarray}$

Figure 8 summarizes our results for the data-driven solution to the Burgers' equation. The top panel of figure 8 compares between the exact dynamics and the recovered behavior. The resulting prediction error is measured at 1.05 × 10−01 in the relative ${{\mathbb{L}}}_{2}$-norm, which is very large. A more detailed assessment of the predicted solution is presented in the bottom panel of figure 8. In particular, we present a comparison between the exact solution and the predicted solution at three different instants t=−7.5, −2.5, 7.5. From the bottom panel, we find that the algorithm can approximately capture the nonlinear behavior of the Burgers' equation. However, the dynamics is hard to accurately resolve from about t=−2.0, which remains unsolved in the current algorithm. It may require more finer-grained spatiotemporal data to be sampled [28], more network layers and hidden neurons [24, 29], even specific random seeds and more advanced network architectures [30-32]. We have made some preliminary experimental attempts, and more detailed analysis will be our future work.

Figure 8.

New window|Download| PPT slide
Figure 8.Exponential initial condition. Top: a solution to the Burgers' equation is compared to the corresponding solution to the learned PDE. The model approximately captures the nonlinear dynamics and reproduces the solution with a relative ${{\mathbb{L}}}_{2}$ error of 1.05 × 10−01. Bottom: comparison between the predicted solution and exact solution at the snapshots. The model training took approximately 12 min.


Then, we consider another exponential initial condition and periodic boundary condition of equation (4) which are given by$\begin{eqnarray}\left\{\begin{array}{l}{u}_{0}(x)=-2(x+5)\exp \left(-{\left(x+5\right)}^{2}\right),\\ u(t,-20)=u(t,20).\end{array}\right.\end{eqnarray}$

Unlike other cases considered in this paper, we just consider the nonlinear dynamics of the equation from t=−3.0 up to t=3.0 here.

Figure 9 summarizes our results for the data-driven solution to the Burgers' equation. The top panel of figure 9 compares between the exact dynamics and the reconstructed spatiotemporal solution. Then the resulting error is measured at 1.97 × 10−02 in the relative ${{\mathbb{L}}}_{2}$-norm. A more detailed assessment of the predicted solution is presented in the bottom panel of figure 9. In particular, we present a comparison between the exact solution and the predicted solution at different moments t=−2.25, −0.75, 2.25. Compared to the previous case, the algorithm can accurately capture the complex nonlinear behavior of the Burgers' equation.

Figure 9.

New window|Download| PPT slide
Figure 9.Another exponential initial condition. Top: a solution to the Burgers' equation is compared to the solution of the learned PDE. The system correctly captures the dynamics and reproduces the solution with a relative ${{\mathbb{L}}}_{2}$ error of 1.97 × 10−02. Bottom: comparison between the predicted and exact solutions at the three temporal snapshots. The training took about half an hour.


5.2. Hyperbolic secant functions

In the following, we will study the hyperbolic secant (sech) function and related functions.

First, we consider a single sech initial condition and periodic boundary condition of equation (4) given by$\begin{eqnarray}\left\{\begin{array}{l}{u}_{0}(x)=3{{\rm{sech}} }^{2}(x+5),\\ u(t,-20)=u(t,20).\end{array}\right.\end{eqnarray}$

Figure 10 summarizes our results for the data-driven solution to the Burgers' equation. The top panel of figure 10 compares between the exact dynamics and the predicted behavior and the resulting error is measured at 3.99 × 10−02 in the relative ${{\mathbb{L}}}_{2}$-norm. A more detailed assessment of the predicted behavior is presented in the bottom panel of figure 10. In particular, we present a comparison between the exact solution and the predicted solution at different instants t=−7.5, −2.5, 7.5. We see that the algorithm can accurately capture the nonlinear behavior of the Burgers' equation. Unlike the case(9), the model accurately resolve the nonlinear dynamics from about t=−2.0.

Figure 10.

New window|Download| PPT slide
Figure 10.Hyperbolic secant initial condition. Top: a solution to the Burgers' equation is compared to the solution to the learned PDE. The system correctly captures the dynamics behavior and reproduces the solution with a relative ${{\mathbb{L}}}_{2}$ error of 3.99 × 10−02. Bottom: comparison between the predicted dynamics and the exact solution at the three temporal snapshots. The model training took approximately 7 min.


Then, we consider a summation of two hyperbolic secant function and periodic boundary condition of equation (4) given by$\begin{eqnarray}\left\{\begin{array}{l}{u}_{0}(x)=2{{\rm{sech}} }^{2}(x-2)+3{{\rm{sech}} }^{2}(x+5),\\ u(t,-20)=u(t,20).\end{array}\right.\end{eqnarray}$

Figure 11 summarizes our results for the data-driven solution to the Burgers' equation. The top panel of figure 11 compares between the exact dynamics and the predicted spatiotemporal solution. The prediction error is measured at 3.83 × 10−02 in the relative ${{\mathbb{L}}}_{2}$-norm. A more detailed assessment of the predicted solution is presented in the bottom panel of figure 11. In particular, we present a comparison between the exact dynamics and the predicted behavior at different time instants t=−7.5, −2.5, 7.5. The algorithm can accurately capture the nonlinear dynamics of the equation.

Figure 11.

New window|Download| PPT slide
Figure 11.Second hyperbolic secant initial condition. Top: a solution to the Burgers' equation is compared to the solution of the learned PDE. The system correctly captures the dynamics and reproduces the solution with a relative ${{\mathbb{L}}}_{2}$ error of 3.83 × 10−02. Bottom: comparison between the predicted solution and the exact solution corresponding to the snapshots. The training took about 48 min.


Furthermore, we also consider another initial condition and periodic boundary condition which are given by$\begin{eqnarray}\left\{\begin{array}{l}{u}_{0}(x)=2{{\rm{sech}} }^{2}(x-2)-3{{\rm{sech}} }^{2}(x+5),\\ u(t,-20)=u(t,20).\end{array}\right.\end{eqnarray}$

Figure 12 summarizes our results for the data-driven solution to the Burgers' equation. The top panel of figure 12 compares between the exact dynamics and the predicted spatiotemporal dynamics and the resulting error is measured at 4.59 × 10−02 in the relative ${{\mathbb{L}}}_{2}$-norm. A more detailed assessment of the predicted dynamics is presented in the bottom panel of figure 12. In particular, we present a comparison between the exact solution and the predicted solution at different time instants t=−7.5, −2.5, 7.5. The algorithm can also accurately capture the nonlinear behavior of the Burgers' equation. In addition, it took much more time than the case(10).

Figure 12.

New window|Download| PPT slide
Figure 12.Another hyperbolic secant initial condition. Top: a solution to the Burgers' equation is compared to the solution of the learned PDE. The system correctly captures the dynamics and reproduces the solution with a relative ${{\mathbb{L}}}_{2}$ error of 4.59 × 10−02. Bottom: comparison between the predicted solution and the exact solution at the snapshots. The training process took approximately 42 min.


6. Remarks and discussion

In this paper, we present a neural network architecture for extracting nonlinear dynamics of PDEs from spatiotemporal data. The framework provides a universal treatment of (1+1)-dimensional second-order nonlinear evolution equations. The resulting method shows a pile of significant results for a diverse collection of initial-boundary conditions, including soliton solutions and other initial data. Relatively speaking, the more complex the initial data, the more time the model training takes. In particular, compared with other initial conditions, the training costs much less time for some soliton solutions perhaps because of the intrinsic structures of solitons.

Note that some of nonlinear evolution PDEs consist of terms like utt, uxt, $\sin (u)$ and some others. It would be very interesting to extend the network framework to incorporate such cases. It will be our future research.

Acknowledgments

We would like to express our sincere thanks to S Y Lou, E G Fan and other members of our discussion group for their valuable comments.


Reference By original order
By published year
By cited within times
By Impact factor

Goodfellow I Bengio Y Courville A 2016 Deep Learning Cambridge, MA MIT Press
[Cited within: 1]

Bongard J Lipson H 2007 Proc. Natl Acad. Sci. USA 104 9943 9948
DOI:10.1073/pnas.0609476104 [Cited within: 1]

Raissi M Perdikaris P Karniadakis G E 2017 J. Comput. Phys. 348 683 693
DOI:10.1016/j.jcp.2017.07.050

Raissi M Perdikaris P Karniadakis G E 2018 SIAM J. Sci. Comput. 40A172 A198
DOI:10.1137/17M1120762 [Cited within: 1]

Raissi M Karniadakis G E 2018 J. Comput. Phys. 357 125 141
DOI:10.1016/j.jcp.2017.11.039 [Cited within: 1]

Cybenko G 1989 Math. Control Signals Syst. 2 303 314
DOI:10.1007/BF02551274 [Cited within: 1]

Hornik K Stinchcombe M White H 1989 Neural Netw. 2 359 366
DOI:10.1016/0893-6080(89)90020-8

Hornik K Stinchcombe M White H 1990 Neural Netw. 3 551 560
DOI:10.1016/0893-6080(90)90005-6

Hornik K 1991 Neural Netw. 4 251 257
DOI:10.1016/0893-6080(91)90009-T

Petersen P Voigtlaender F 2018 Neural Netw. 108 296 330
DOI:10.1016/j.neunet.2018.08.019 [Cited within: 1]

Baydin A G Pearlmutter B A Radul A A Siskind J M 2018 J. Mach. Learn. Res. 18 1 43
[Cited within: 1]

Rudy S H Brunton S L Proctor J L Kutz J N 2017 Sci. Adv. 3e1602614
DOI:10.1126/sciadv.1602614 [Cited within: 1]

Murray J D 1993 Mathematical Biology Berlin Springer
[Cited within: 1]

Lagaris I E Likas A Fotiadis D I 1998 IEEE Trans. Neural Netw. 9 987 1000
DOI:10.1109/72.712178 [Cited within: 2]

Yadav N Yadav A Kumar M 2015 An Introduction to Neural Network Methods for Differential Equations Berlin Springer


Sirignano J Spiliopoulos K 2018 J. Comput. Phys. 375 1339 1364
DOI:10.1016/j.jcp.2018.08.029

Han J Jentzen A Weinan E 2018 Proc. Natl Acad. Sci. USA 115 8505 8510
DOI:10.1073/pnas.1718942115

Bar-Sinai Y Hoyer S Hickey J Brenner M P 2019 Proc. Natl Acad. Sci. USA 116 15344 15349
DOI:10.1073/pnas.1814058116 [Cited within: 1]

Abadi M et al. 201612th USENIX Symp. on Operating Systems Design and Implementation 16 265 283
[Cited within: 1]

Kingma D P Ba J 2015Int. Conf. on Learning Representations (ICLR)
[Cited within: 1]

Choromanska A Henaff M Mathieu M Arous G B Lecun Y 2015Proc. 18th Int. Conf. on Artificial Intelligence and Statistics (AISTATS) vol 38
[Cited within: 1]

Driscoll T A Hale N Trefethen L N 2014 Chebfun Guide Oxford Pafnuty Publications
[Cited within: 1]

Stein M L 1987 Technometrics 29 143 151
DOI:10.1080/00401706.1987.10488205 [Cited within: 1]

Raghu M Poole B Kleinberg J Ganguli S Sohl-Dickstein J 2017Proc. 34th Int. Conf. on Machine Learning (ICML)
[Cited within: 1]

Wang S Tang X Lou S 2004 Chaos Solitons Fractals 21 231 239
DOI:10.1016/j.chaos.2003.10.014 [Cited within: 2]

Raissi M 2018 J. Mach. Learn. Res. 19 932 955
[Cited within: 3]

Raissi M Perdikaris P Karniadakis G E 2019 J. Comput. Phys. 378 686 707
DOI:10.1016/j.jcp.2018.10.045

Lu L Meng X Mao Z Karniadakis G E 2019 DeepXDE: a deep learning library for solving differential equations. (arXiv:1907.04502)
[Cited within: 2]

Lin H W Tegmark M Rolnick D 2017 J. Stat. Phys. 168 1223 1247
DOI:10.1007/s10955-017-1836-5 [Cited within: 1]

Mallat S 2016 Phil. Trans. R. Soc. A 374 20150203
DOI:10.1098/rsta.2015.0203 [Cited within: 1]

Hagge T Stinis P Yeung E Tartakovsky A M 2017 Solving differential equations with unknown constitutive relations as recurrent neural networks (arXiv:1710.02242)


Michoski C Milosavljević M Oliver T Hatch D 2019 Solving irregular and data-enriched differential equations using deep neural networks (arXiv:1905.04351)
[Cited within: 2]

相关话题/Solving secondorder nonlinear